This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A combinatorial integration on the Cantor dust

Takashi MARUYAMA Department of Engineering, Stanford University, 353 Jane Stanford Way Stanford CA 94305, USA [email protected], [email protected]  and  Tatsuki SETO General Education and Research Center, Meiji Pharmaceutical University, 2-522-1 Noshio, Kiyose-shi, Tokyo, Japan [email protected]
Abstract.

In this paper, we generalize the Cantor function to 22-dimensional cubes and construct a cyclic 22-cocycle on the Cantor dust. This cocycle is non-trivial on the pullback of the smooth functions on the 22-dimensional torus with the generalized Cantor function while it vanishes on the Lipschitz functions on the Cantor dust. The cocycle is calculated through the integration of 22-forms on the torus by using a combinatorial Fredholm module.

Key words and phrases:
Fredholm module, Cantor dust, cyclic cocycle
2020 Mathematics Subject Classification:
Primary 46L87; Secondary 28A80.

Introduction

Cyclic cohomology for algebras [2] is a fundamental tool to study noncommutative geometry. One of its application is the study of fractal sets to which powerful tools such as de Rham homology (on smooth manifolds) cannot be applied. KK-homology, one of other homotopy invariant theories, of fractal sets is also studied extensively. For some class of fractal sets such as the Cantor set and the Cantor dust, the K0K^{0} homology groups are isomorphic to \displaystyle\prod^{\infty}\mathbb{Z}, which is known as the Baer-Specker group [1]. One feature of the Baer-Specker group is that the group does not admit basis. This feature makes the study of K0K^{0} groups for such fractal sets somewhat intractable because local topological features of the spaces cannot induce local algebraic structures. Cyclic cohomology on the one hand is expected to be favorable in these cases because it can be characterized as a “linearization” of KK-homology through the Chern character.

There is a study [4] by H. Moriyoshi and T. Natsume which presented a variant of the Riemann-Stieltjes integration on the middle third Cantor set. Let CS=n=0InCS=\displaystyle\bigcap_{n=0}^{\infty}I_{n} be the middle third Cantor set, where I0=[0,1]I_{0}=[0,1], I1=[0,13]I_{1}=\left[0,\dfrac{1}{3}\right], I2=[0,132][232,332][632,732][832,1]I_{2}=\left[0,\dfrac{1}{3^{2}}\right]\cup\left[\dfrac{2}{3^{2}},\dfrac{3}{3^{2}}\right]\cup\left[\dfrac{6}{3^{2}},\dfrac{7}{3^{2}}\right]\cup\left[\dfrac{8}{3^{2}},1\right], …. We also denote by (Hn,Fn)(H_{n},F_{n}) the Fredholm module on InI_{n} which is defined by the direct sum of Connes’ Fredholm module on intervals. In [4], a new class of algebra is introduced: an algebra 𝒫=cBV(S1)\mathcal{P}=c^{\ast}BV(S^{1}) defined by the pull-back of the bounded variation class on a unit circle BV(S1)BV(S^{1}) with the canonical Cantor function cc. They defined a functional ψ(f,g)=limnψn(f,g)=limnTr(ϵf[Fn,g]Fn)\psi(f,g)=\displaystyle\lim_{n\to\infty}\psi_{n}(f,g)=\lim_{n\to\infty}\mathrm{Tr}(\epsilon f[F_{n},g]F_{n}) on 𝒫\mathcal{P}. The existence of the limit can be proved by showing that the cocycle is reduced to the Riemann-Stieltjes integration. A key ingredient for the proof is the Cantor function (and its intermediate functions that converge into the Cantor function); the function is used to map the iterated function system of Cantor sets onto 22-fold subdivisions of the unit interval on which the Riemann-Stieltjes integration is defined. Because the Cantor function remains surjective onto the unit interval when the function is restricted to the Cantor set, the iterated function system for the Cantor set may be seen as a variant of subdivisions on II.

In the present paper, by following the idea mentioned above, we construct a new cyclic 22-cocycle on the Cantor dust CDCD and show that the cocycle is not trivial. In order to construct the cocycle, we define a sequence of functionals ϕn\phi_{n} (see Definition 3.2), that is a generalization of a functional ψn\psi_{n} on CSCS, by using a combinatorial Fredholm module on squares defined by the authors [3] and use the product of the Cantor function. The function induces the map 𝒄\boldsymbol{c} between the Cantor dust and 22-torus 𝕋2\mathbb{T}^{2}:

Theorem (see Theorem 3.4).

Let C(𝕋2)C^{\infty}(\mathbb{T}^{2}) be the smooth functions on the torus 𝕋2\mathbb{T}^{2}. For f=𝐜(f~),g=𝐜(g~),h=𝐜(h~)𝐜C(𝕋2)f=\mbox{$\boldsymbol{c}$}^{*}(\tilde{f}),g=\mbox{$\boldsymbol{c}$}^{*}(\tilde{g}),h=\mbox{$\boldsymbol{c}$}^{*}(\tilde{h})\in\mbox{$\boldsymbol{c}$}^{*}C^{\infty}(\mathbb{T}^{2}), we have

limnϕn(f,g,h)=2𝕋2f~𝑑g~dh~.\lim_{n\to\infty}\phi_{n}(f,g,h)=2\int_{\mathbb{T}^{2}}\tilde{f}d\tilde{g}\wedge d\tilde{h}.

Moreover, the limit only depends on f,g,h𝐜C(𝕋2)f,g,h\in\mbox{$\boldsymbol{c}$}^{*}C^{\infty}(\mathbb{T}^{2}). Therefore, the functional ϕ\phi defined by the limit is a cyclic 22-cocycle on 𝐜C(𝕋2)\mbox{$\boldsymbol{c}$}^{*}C^{\infty}(\mathbb{T}^{2}).

We can take the value ϕn(f,g,h)\phi_{n}(f,g,h) for Lipschitz functions f,g,hCLip(CD)f,g,h\in C^{\mathrm{Lip}}(CD) by the definition of ϕn\phi_{n}. However, we have limnϕn(f,g,h)=0\displaystyle\lim_{n\to\infty}\phi_{n}(f,g,h)=0 for any f,g,hCLip(CD)f,g,h\in C^{\mathrm{Lip}}(CD) (see Proposition 3.3). Thus the algebra 𝒄C(𝕋2)\mbox{$\boldsymbol{c}$}^{*}C^{\infty}(\mathbb{T}^{2}) is far different from CLip(CD)C^{\mathrm{Lip}}(CD).

All the main results to be shown also hold for the Sierpinski carpet. 𝕋2\mathbb{T}^{2} can be also replaced with a 22-dimensional sphere. In order to generalize our main theorem to general dimension, we need a general method to prove the existence of a cyclic cocycle. We will leave the study for future work.

1. A combinatorial Fredholm module on squares

In this section, we review a combinatorial Fredholm module on squares constructed by the authors [3]. Let γ2\gamma\subset\mathbb{R}^{2} be a square of dimension 22 and V={v0,v1,v2,v3}V=\{v_{0},\ v_{1},\ v_{2},\ v_{3}\} the set of vertices of γ\gamma; see the following figure for the numbering of the vertices.

v3\textstyle{v_{3}}v2\textstyle{v_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}v0\textstyle{v_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}v1\textstyle{v_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}
Figure 1. Numbering of the vertices

Set V0={v0,v2}V_{0}=\{v_{0},\ v_{2}\} and V1={v1,v3}V_{1}=\{v_{1},\ v_{3}\}, so we have V=V0V1V=V_{0}\cup V_{1}. Set

+\displaystyle\mathcal{H}^{+} =2(V0)=2(v0)2(v2),\displaystyle=\ell^{2}(V_{0})=\ell^{2}(v_{0})\oplus\ell^{2}(v_{2}), \displaystyle\mathcal{H}^{-} =2(V1)=2(v1)2(v3)\displaystyle=\ell^{2}(V_{1})=\ell^{2}(v_{1})\oplus\ell^{2}(v_{3})

and =+\mathcal{H}=\mathcal{H}^{+}\oplus\mathcal{H}^{-}. The vector space (4)\mathcal{H}(\cong\mathbb{C}^{4}) is a Hilbert space of dimension 44 with an inner product

f,g=i=03f(vi)g(vi)¯.\langle f,g\rangle=\sum_{i=0}^{3}f(v_{i})\overline{g(v_{i})}.

We assume that \mathcal{H} is 2\mathbb{Z}_{2}-graded with the grading ϵ=±1\epsilon=\pm 1 on ±\mathcal{H}^{\pm}, respectively. The CC^{\ast}-algebra C(V)C(V) of continuous functions on VV acts on \mathcal{H} by multiplication:

ρ(f)=(f(v0)f(v2))(f(v1)f(v3)).\rho(f)=(f(v_{0})\oplus f(v_{2}))\oplus(f(v_{1})\oplus f(v_{3})).

Set U=12[1111]U=\dfrac{1}{\sqrt{2}}\begin{bmatrix}1&-1\\ 1&1\end{bmatrix} and F=[UU]=12[11111111]F=\begin{bmatrix}&U^{\ast}\\ U&\end{bmatrix}=\dfrac{1}{\sqrt{2}}\begin{bmatrix}&&1&1\\ &&-1&1\\ 1&-1&&\\ 1&1&&\end{bmatrix}. Then FF is a bounded operator on \mathcal{H} and we have Fϵ+ϵF=OF\epsilon+\epsilon F=O. So (,F)(\mathcal{H},F) is the Fredholm module on C(V)C(V).

Set I=[0,1]×[0,1]I=[0,1]\times[0,1] and let fs:IIf_{s}:I\to I (s=1,,Ns=1,\dots,N) be similitudes. Denote by

rs=fs(x)fs(y)nxyn(<1)(xy)r_{s}=\dfrac{\|f_{s}(x)-f_{s}(y)\|_{\mathbb{R}^{n}}}{\|x-y\|_{\mathbb{R}^{n}}}\;\,(<1)\quad(x\neq y)

the similarity ratio of fsf_{s}. An iterated function system (IFS) (I,S={1,,N},{fs}sS)(I,S=\{1,\dots,N\},\{f_{s}\}_{s\in S}) defines the unique non-empty compact set K=K(γn,S={1,,N},{fs}sS)K=K(\gamma_{n},S=\{1,\dots,N\},\{f_{s}\}_{s\in S}) called the self-similar set such that K=s=1Nfs(K)K=\bigcup_{s=1}^{N}f_{s}(K). Denote by dimS(K)\dim_{S}(K) the similarity dimension of KK, that is, the number ss that satisfies

s=1Nrss=1.\sum_{s=1}^{N}r_{s}^{s}=1.

If an IFS (I,S,{fs}sS)(I,S,\{f_{s}\}_{s\in S}) satisfies the open set condition, we have dimH(K)=dimS(K)\dim_{H}(K)=\dim_{S}(K), where we denote by dimH(K)\dim_{H}(K) the Hausdorff dimension of KK.

Set f𝒔=fs1fsjf_{\mbox{$\boldsymbol{s}$}}=f_{s_{1}}\circ\dots\circ f_{s_{j}} for 𝒔=(s1,,sj)S=j=0S×j\mbox{$\boldsymbol{s}$}=(s_{1},\dots,s_{j})\in S^{\infty}=\bigcup_{j=0}^{\infty}S^{\times j} and f=idf_{\emptyset}=\mathrm{id}. For the simplicity, we will denote by ii the vertex f𝒔(vi)f_{\mbox{$\boldsymbol{s}$}}(v_{i}) of a square f𝒔(I)f_{\mbox{$\boldsymbol{s}$}}(I). We also denote by V𝒔V_{\mbox{$\boldsymbol{s}$}} the vertices of a square f𝒔(I)f_{\mbox{$\boldsymbol{s}$}}(I). Denote by e𝒔e_{\mbox{$\boldsymbol{s}$}} the length of edge of f𝒔(I)f_{\mbox{$\boldsymbol{s}$}}(I), which equals k=1jrsk\prod_{k=1}^{j}r_{s_{k}}. As introduced above, we set the Hilbert space 𝒔=2(V𝒔)\mathcal{H}_{\mbox{$\boldsymbol{s}$}}=\ell^{2}(V_{\mbox{$\boldsymbol{s}$}}) on f𝒔(I)f_{\mbox{$\boldsymbol{s}$}}(I) of the length e𝒔e_{\mbox{$\boldsymbol{s}$}}, which splits the positive part 𝒔+\mathcal{H}_{\mbox{$\boldsymbol{s}$}}^{+} and the negative part 𝒔\mathcal{H}_{\mbox{$\boldsymbol{s}$}}^{-}. Taking direct sum on squares, we set as follows:

n=j=1n𝒔S×j𝒔,Fn=j=1n𝒔S×jF.\mathcal{H}_{n}=\bigoplus_{j=1}^{n}\bigoplus_{\mbox{$\boldsymbol{s}$}\in S^{\times j}}\mathcal{H}_{\mbox{$\boldsymbol{s}$}},\quad F_{n}=\bigoplus_{j=1}^{n}\bigoplus_{\mbox{$\boldsymbol{s}$}\in S^{\times j}}F.

The CC^{\ast}-algebra C(K)C(K) acts on n\mathcal{H}_{n} by multiplication:

ρn(f)=j=1n𝒔S×jρ(f).\rho_{n}(f)=\bigoplus_{j=1}^{n}\bigoplus_{\mbox{$\boldsymbol{s}$}\in S^{\times j}}\rho(f).

2. Review on the Cantor dust

We introduce an IFS of the Cantor dust and 22-dimensional analogue of the Cantor function defined on 11-dimensional interval. The IFS of the Cantor dust is our main interest in the paper. The analogue of the Cantor function plays a crucial role in construction of cyclic cocycle in Section 3. Let I=[0,1]×[0,1]I=[0,1]\times[0,1]. The IFS of the Cantor dust is defined as a set of functions {fs:I2}s=1,2,3,4\{f_{s}:I\rightarrow\mathbb{R}^{2}\}_{s=1,2,3,4} described as follows:

f1(x)=13x,f2(x)=13x+13[02],f3(x)=13x+13[20],f4(x)=13x+13[22].\displaystyle f_{1}(x)=\frac{1}{3}x,\ f_{2}(x)=\frac{1}{3}x+\dfrac{1}{3}\begin{bmatrix}0\\ 2\end{bmatrix},\ f_{3}(x)=\frac{1}{3}x+\dfrac{1}{3}\begin{bmatrix}2\\ 0\end{bmatrix},\ f_{4}(x)=\frac{1}{3}x+\dfrac{1}{3}\begin{bmatrix}2\\ 2\end{bmatrix}.

This IFS induces a unique non-empty compact set CDCD in II such that CD=i=14fs(CD)\displaystyle CD=\bigcup_{i=1}^{4}f_{s}(CD) holds; CDCD is called the Cantor dust.

Let c0=x:[0,1]c_{0}=x:[0,1]\rightarrow\mathbb{R}. We define in an inductive manner a sequence of \mathbb{R}-valued continuous functions defined on the unit interval {cn}n\{c_{n}\}_{n\in\mathbb{N}} as follows:

cn+1(x)={12cn(3x),0x13,12,13x23,12cn(3x2)+12,23x1.c_{n+1}(x)=\begin{cases}\frac{1}{2}c_{n}(3x),&0\leq x\leq\frac{1}{3},\\ \frac{1}{2},&\frac{1}{3}\leq x\leq\frac{2}{3},\\ \frac{1}{2}c_{n}(3x-2)+\frac{1}{2},&\frac{2}{3}\leq x\leq 1.\end{cases}

The limit of {cn}n\{c_{n}\}_{n\in\mathbb{N}} exists and is called the Cantor function. We then generalize the construction to 22-dimensional case by taking the pair of the two identical sequences:

𝒄n=(cn,cn):I2.\mbox{$\boldsymbol{c}$}_{n}=(c_{n},c_{n}):I\rightarrow\mathbb{R}^{2}.

The limit of the sequence {𝒄n}\{\mbox{$\boldsymbol{c}$}_{n}\} also exists and equals the product of the Cantor functions. We call the limit limn𝒄n\displaystyle\lim_{n\to\infty}\mbox{$\boldsymbol{c}$}_{n} 22-dimensional Cantor dust function. We note that the construction of the 22-dimensional Cantor dust function can be generalized to higher dimensional case.

3. A combinatorial integration

3.1. Approximating combinatorial integration

We apply the construction of the combinatorial Fredholm module to the IFS of the Cantor dust (I,S={1,2,3,4},{fs}sS)(I,S=\{1,2,3,4\},\{f_{s}\}_{s\in S}) defined in Section 2. In order to construct our cyclic cocycle, we need the multiplication operator ρ(f)\rho(f) and the commutator [F,f][F,f] on 𝒔\mathcal{H}_{\mbox{$\boldsymbol{s}$}} for fC(CD)f\in C(CD). ρ(f)\rho(f) is even, so we can express ρ(f)=[f+f]\rho(f)=\begin{bmatrix}f^{+}&\\ &f^{-}\end{bmatrix}. We can write down the commutator [F,f][F,f] as follows:

[F,f]=12[(f(0)f(1))(f(0)f(3))f(2)f(1)(f(2)f(3))f(0)f(1)(f(2)f(1))f(0)f(3)f(2)f(3)].[F,f]=\frac{1}{\sqrt{2}}\begin{bmatrix}&&-\left(f(0)-f(1)\right)&-\left(f(0)-f(3)\right)\\ &&f(2)-f(1)&-\left(f(2)-f(3)\right)\\ f(0)-f(1)&-\left(f(2)-f(1)\right)&&\\ f(0)-f(3)&f(2)-f(3)&&\end{bmatrix}.

Here, for the simplicity, we denote i=vii=v_{i} the vertices on the squares V𝒔V_{\mbox{$\boldsymbol{s}$}}. We denote the upper right 2×22\times 2 block of [F,f][F,f] by dfd^{-}f and the lower left by d+fd^{+}f.

We now construct a sequence of operators on 𝒔S×n𝒔\displaystyle\bigoplus_{\mbox{$\boldsymbol{s}$}\in S^{\times n}}\mathcal{H}_{\mbox{$\boldsymbol{s}$}} that will give rise to a cyclic cocycle. For f,g,hC(K)f,g,h\in C(K), we have

f[F,g][F,h]\displaystyle f[F,g][F,h] =[f+f][dgd+g][dhd+h]\displaystyle=\begin{bmatrix}f^{+}&\\ &f^{-}\end{bmatrix}\begin{bmatrix}&d^{-}g\\ d^{+}g&\end{bmatrix}\begin{bmatrix}&d^{-}h\\ d^{+}h&\end{bmatrix}
=[f+dgfd+g][dhd+h]=[f+dgd+hfd+gdh].\displaystyle=\begin{bmatrix}&f^{+}d^{-}g\\ f^{-}d^{+}g&\end{bmatrix}\begin{bmatrix}&d^{-}h\\ d^{+}h&\end{bmatrix}=\begin{bmatrix}f^{+}d^{-}gd^{+}h&\\ &f^{-}d^{+}gd^{-}h\end{bmatrix}.

By setting fi,j=f(j)f(i)f_{i,j}=f(j)-f(i), the two diagonal 2×22\times 2 blocks of f[F,g][F,h]f[F,g][F,h] can be expressed as

f+dgd+h\displaystyle f^{+}d^{-}gd^{+}h =12[f(0)f(2)][g0,1g0,3g2,1g2,3][h0,1h2,1h0,3h2,3]\displaystyle=\frac{1}{2}\begin{bmatrix}f(0)&\\ &f(2)\end{bmatrix}\begin{bmatrix}g_{0,1}&g_{0,3}\\ -g_{2,1}&g_{2,3}\end{bmatrix}\begin{bmatrix}-h_{0,1}&h_{2,1}\\ -h_{0,3}&-h_{2,3}\end{bmatrix}
=12[f(0)g0,1f(0)g0,3f(2)g2,1f(2)g2,3][h0,1h2,1h0,3h2,3]\displaystyle=\frac{1}{2}\begin{bmatrix}f(0)g_{0,1}&f(0)g_{0,3}\\ -f(2)g_{2,1}&f(2)g_{2,3}\end{bmatrix}\begin{bmatrix}-h_{0,1}&h_{2,1}\\ -h_{0,3}&-h_{2,3}\end{bmatrix}
=12[f(0)(g0,1h0,1+g0,3h0,3)f(0)(g0,1h2,1g0,3h2,3)f(2)(g2,1h0,1g2,3h0,3)f(2)(g2,1h2,1+g2,3h2,3)]\displaystyle=\frac{1}{2}\begin{bmatrix}-f(0)(g_{0,1}h_{0,1}+g_{0,3}h_{0,3})&f(0)(g_{0,1}h_{2,1}-g_{0,3}h_{2,3})\\ f(2)(g_{2,1}h_{0,1}-g_{2,3}h_{0,3})&-f(2)(g_{2,1}h_{2,1}+g_{2,3}h_{2,3})\end{bmatrix}
=12[f(0)(g0,1h1,0+g0,3h3,0)f(0)(g0,1h1,2g0,3h3,2)f(2)(g2,3h3,0g2,1h1,0)f(2)(g2,1h1,2+g2,3h3,2)]\displaystyle=\frac{1}{2}\begin{bmatrix}f(0)(g_{0,1}h_{1,0}+g_{0,3}h_{3,0})&-f(0)(g_{0,1}h_{1,2}-g_{0,3}h_{3,2})\\ f(2)(g_{2,3}h_{3,0}-g_{2,1}h_{1,0})&f(2)(g_{2,1}h_{1,2}+g_{2,3}h_{3,2})\end{bmatrix}

and

fd+gdh\displaystyle f^{-}d^{+}gd^{-}h =12[f(1)f(3)][g0,1g2,1g0,3g2,3][h0,1h0,3h2,1h2,3]\displaystyle=\frac{1}{2}\begin{bmatrix}f(1)&\\ &f(3)\end{bmatrix}\begin{bmatrix}-g_{0,1}&g_{2,1}\\ -g_{0,3}&-g_{2,3}\end{bmatrix}\begin{bmatrix}h_{0,1}&h_{0,3}\\ -h_{2,1}&h_{2,3}\end{bmatrix}
=12[f(1)g0,1f(1)g2,1f(3)g0,3f(3)g2,3][h0,1h0,3h2,1h2,3]\displaystyle=\frac{1}{2}\begin{bmatrix}-f(1)g_{0,1}&f(1)g_{2,1}\\ -f(3)g_{0,3}&-f(3)g_{2,3}\end{bmatrix}\begin{bmatrix}h_{0,1}&h_{0,3}\\ -h_{2,1}&h_{2,3}\end{bmatrix}
=12[f(1)(g0,1h0,1g2,1h2,1)f(1)(g0,1h0,3+g2,1h2,3)f(3)(g0,3h0,1+g2,3h2,1)f(3)(g0,3h0,3g2,3h2,3)]\displaystyle=\frac{1}{2}\begin{bmatrix}f(1)(-g_{0,1}h_{0,1}-g_{2,1}h_{2,1})&f(1)(-g_{0,1}h_{0,3}+g_{2,1}h_{2,3})\\ f(3)(-g_{0,3}h_{0,1}+g_{2,3}h_{2,1})&f(3)(-g_{0,3}h_{0,3}-g_{2,3}h_{2,3})\end{bmatrix}
=12[f(1)(g1,0h0,1+g1,2h2,1)f(1)(g1,0h0,3g1,2h2,3)f(3)(g3,2h2,1g3,0h0,1)f(3)(g3,0h0,3+g3,2h2,3)].\displaystyle=\frac{1}{2}\begin{bmatrix}f(1)(g_{1,0}h_{0,1}+g_{1,2}h_{2,1})&f(1)(g_{1,0}h_{0,3}-g_{1,2}h_{2,3})\\ -f(3)(g_{3,2}h_{2,1}-g_{3,0}h_{0,1})&f(3)(g_{3,0}h_{0,3}+g_{3,2}h_{2,3})\end{bmatrix}.

Here, we define

M=2e2[F,x][F,y]=[1111].M=-\frac{2}{e^{2}}[F,x][F,y]=\begin{bmatrix}&1&&\\ -1&&&\\ &&&1\\ &&-1&\end{bmatrix}.

MM can be expressed as NNN\oplus N by denoting the off-diagonal 2×22\times 2 matrices by NN. The following lemma indicates that the trace of an operator f[F,g][F,h]Mf[F,g][F,h]M gives rise to a discretized version of integration on a square.

Lemma 3.1.

For any nn\in\mathbb{N} and 𝐬S×n\mbox{$\boldsymbol{s}$}\in S^{\times n}, we have

2Tr(f[F,g][F,h]M)=\displaystyle 2\cdot\operatorname{Tr}(f[F,g][F,h]M)= f(0)(g0,1h1,2g0,3h3,2)+f(2)(g2,3h3,0g2,1h1,0)\displaystyle f(0)(g_{0,1}h_{1,2}-g_{0,3}h_{3,2})+f(2)(g_{2,3}h_{3,0}-g_{2,1}h_{1,0})
f(1)(g1,0h0,3g1,2h2,3)f(3)(g3,2h2,1g3,0h0,1).\displaystyle-f(1)(g_{1,0}h_{0,3}-g_{1,2}h_{2,3})-f(3)(g_{3,2}h_{2,1}-g_{3,0}h_{0,1}).
Proof.

The proof follows straightforward calculation. By multiplying MM with f[F,g][F,h]f[F,g][F,h], we have

f[F,g][F,h]M\displaystyle f[F,g][F,h]M =[f+dgd+hfd+gdh][NN]\displaystyle=\begin{bmatrix}f^{+}d^{-}gd^{+}h&\\ &-f^{-}d^{+}gd^{-}h\end{bmatrix}\begin{bmatrix}N&\\ &N\end{bmatrix}
=[f+dgd+hNfd+gdhN].\displaystyle=\begin{bmatrix}f^{+}d^{-}gd^{+}hN&\\ &f^{-}d^{+}gd^{-}h\;N\end{bmatrix}.

Then we have

f+dgd+hN\displaystyle f^{+}d^{-}gd^{+}hN =12[f(0)(g0,1h1,0+g0,3h3,0)f(0)(g0,1h1,2g0,3h3,2)f(2)(g2,3h3,0g2,1h1,0)f(2)(g2,1h1,2+g2,3h3,2)][11]\displaystyle=\frac{1}{2}\begin{bmatrix}f(0)(g_{0,1}h_{1,0}+g_{0,3}h_{3,0})&-f(0)(g_{0,1}h_{1,2}-g_{0,3}h_{3,2})\\ f(2)(g_{2,3}h_{3,0}-g_{2,1}h_{1,0})&f(2)(g_{2,1}h_{1,2}+g_{2,3}h_{3,2})\end{bmatrix}\begin{bmatrix}&1\\ -1&\end{bmatrix}
=12[f(0)(g0,1h1,2g0,3h3,2)f(0)(g0,1h1,0+g0,3h3,0)f(2)(g2,1h1,2+g2,3h3,2)f(2)(g2,3h3,0g2,1h1,0)]\displaystyle=\frac{1}{2}\begin{bmatrix}f(0)(g_{0,1}h_{1,2}-g_{0,3}h_{3,2})&f(0)(g_{0,1}h_{1,0}+g_{0,3}h_{3,0})\\ -f(2)(g_{2,1}h_{1,2}+g_{2,3}h_{3,2})&f(2)(g_{2,3}h_{3,0}-g_{2,1}h_{1,0})\end{bmatrix}

and

fd+gdhN\displaystyle f^{-}d^{+}gd^{-}h\;N =12[f(1)(g1,0h0,1+g1,2h2,1)f(1)(g1,0h0,3g1,2h2,3)f(3)(g3,2h2,1g3,0h0,1)f(3)(g3,0h0,3+g3,2h2,3)][11]\displaystyle=\frac{1}{2}\begin{bmatrix}f(1)(g_{1,0}h_{0,1}+g_{1,2}h_{2,1})&f(1)(g_{1,0}h_{0,3}-g_{1,2}h_{2,3})\\ -f(3)(g_{3,2}h_{2,1}-g_{3,0}h_{0,1})&f(3)(g_{3,0}h_{0,3}+g_{3,2}h_{2,3})\end{bmatrix}\begin{bmatrix}&1\\ -1&\end{bmatrix}
=12[f(1)(g1,0h0,3g1,2h2,3)f(1)(g1,0h0,1+g1,2h2,1)f(3)(g3,0h0,3+g3,2h2,3)f(3)(g3,2h2,1g3,0h0,1)].\displaystyle=\frac{1}{2}\begin{bmatrix}-f(1)(g_{1,0}h_{0,3}-g_{1,2}h_{2,3})&f(1)(g_{1,0}h_{0,1}+g_{1,2}h_{2,1})\\ -f(3)(g_{3,0}h_{0,3}+g_{3,2}h_{2,3})&-f(3)(g_{3,2}h_{2,1}-g_{3,0}h_{0,1})\end{bmatrix}.

Therefore the trace of f[F,g][F,h]Mf[F,g][F,h]M is given by the lemma. ∎

When g=xg=x and h=yh=y are the coordinate functions of 2\mathbb{R}^{2}, respectively, each term of the right hand side of Lemma 3.1 is nothing but the summand of the Riemannian sum of ff. So the sum of the trace in the left hand side can be regard as an analogue of the Riemannian sum on the Cantor dust.

Definition 3.2.

For any nn\in\mathbb{N} and fghC(CD)C(CD)C(CD)f\otimes g\otimes h\in C(CD)\otimes C(CD)\otimes C(CD), define a map

ϕn(f,g,h)=𝒔S×nTr(f[F,g][F,h]M).\phi_{n}(f,g,h)=\sum_{\mbox{$\boldsymbol{s}$}\in S^{\times n}}\operatorname{Tr}(f[F,g][F,h]M).

We call the sequence of the maps {ϕn}\{\phi_{n}\} the approximating combinatorial integration.

Proposition 3.3.

For any functions fC(CD)f\in C(CD) and g,hCLip(CD)g,h\in C^{\mathrm{Lip}}(CD), the sequence {ϕn(f,g,h)}\{\phi_{n}(f,g,h)\} converges to 0.

Proof.

By Lemma 3.1, we have

|ϕn(f,g,h)|\displaystyle\bigl{|}\phi_{n}(f,g,h)\bigr{|} 𝒔S×n(|f(0)(g0,1h1,2g0,3h3,2)|+|f(2)(g2,3h3,0g2,1h1,0)|\displaystyle\leq\sum_{\mbox{$\boldsymbol{s}$}\in S^{\times n}}\Bigl{(}\bigl{|}f(0)(g_{0,1}h_{1,2}-g_{0,3}h_{3,2})\bigr{|}+\bigl{|}f(2)(g_{2,3}h_{3,0}-g_{2,1}h_{1,0})\bigr{|}
+|f(1)(g1,0h0,3g1,2h2,3)|+|f(3)(g3,2h2,1g3,0h0,1)|)\displaystyle\hskip 42.67912pt+\bigl{|}f(1)(g_{1,0}h_{0,3}-g_{1,2}h_{2,3})\bigr{|}+\bigl{|}f(3)(g_{3,2}h_{2,1}-g_{3,0}h_{0,1})\bigr{|}\Bigr{)}
2Lip(g)Lip(h)19n𝒔S×n(|f(0)|+|f(1)|+|f(2)|+|f(3)|)\displaystyle\leq 2\mathrm{Lip}(g)\mathrm{Lip}(h)\dfrac{1}{9^{n}}\sum_{\mbox{$\boldsymbol{s}$}\in S^{\times n}}\left(|f(0)|+|f(1)|+|f(2)|+|f(3)|\right)
8fLip(g)Lip(h)4n9n\displaystyle\leq 8\|f\|\mathrm{Lip}(g)\mathrm{Lip}(h)\dfrac{4^{n}}{9^{n}}

for any nn\in\mathbb{N}, fC(CD)f\in C(CD), and g,hCLip(CD)g,h\in C^{\mathrm{Lip}}(CD). Here, f\|f\| means the supnorm of ff and Lip(g)\mathrm{Lip}(g) means the Lipschitz constant of gCLip(CD)g\in C^{\mathrm{Lip}}(CD). Therefore, we have ϕn(f,g,h)0\phi_{n}(f,g,h)\to 0 (nn\to\infty). ∎

By Proposition 3.3, our approximating combinatorial integration does not restore rich structure on the Lipschitz functions. So we need anothor class of functions on the Cantor dust.

3.2. Non-triviality of a combinatorial integration

Let CDn=𝒔S×nf𝒔(I)CD_{n}=\bigcup_{\mbox{$\boldsymbol{s}$}\in S^{\times n}}f_{\mbox{$\boldsymbol{s}$}}(I) be a space obtained by applying the nn-times composition of s=14fs\displaystyle\bigcup_{s=1}^{4}f_{s} to II. CDnCD_{n} has 4n4^{n} distinct squares {γin}1i4n\{\gamma^{n}_{i}\}_{1\leq i\leq 4^{n}} with length of edge 13n\displaystyle\frac{1}{3^{n}}. For every nn\in\mathbb{N}, the restriction of 𝒄n\mbox{$\boldsymbol{c}$}_{n} to CDnCD_{n} is still surjective. The image of {γin}1i4n\{\gamma^{n}_{i}\}_{1\leq i\leq 4^{n}} by 𝒄n\mbox{$\boldsymbol{c}$}_{n} is then a subdivision of II each of whose cell is a square with length 12n\displaystyle\frac{1}{2^{n}}. Therefore, the maximum length across all the cells tends to 0 as nn\to\infty.

We introduce an equivalence relation \sim on I\partial I: (a,0)(a,1)(a,0)\sim(a,1) and (0,b)(1,b)(0,b)\sim(1,b). The quotient space turns out to be the torus 𝕋2=I/\mathbb{T}^{2}=I/\sim. The map 𝒄n\mbox{$\boldsymbol{c}$}_{n} induces a surjective map CDn𝕋2CD_{n}\to\mathbb{T}^{2}. We denote the map by the same letter 𝒄n\mbox{$\boldsymbol{c}$}_{n}. Similarly, we can also construct a continuous function 𝒄:CD𝕋2\mbox{$\boldsymbol{c}$}:CD\rightarrow\mathbb{T}^{2} by the restriction of the 2-dimensional Cantor dust function to CDCD.

The following theorem shows that the limit of approximating combinatorial integration with nontrivial value exists for some class of function on CDCD.

Theorem 3.4.

Let C(𝕋2)C^{\infty}(\mathbb{T}^{2}) be the smooth functions on the torus 𝕋2\mathbb{T}^{2}. For f=𝐜(f~),g=𝐜(g~),h=𝐜(h~)𝐜C(𝕋2)f=\mbox{$\boldsymbol{c}$}^{*}(\tilde{f}),g=\mbox{$\boldsymbol{c}$}^{*}(\tilde{g}),h=\mbox{$\boldsymbol{c}$}^{*}(\tilde{h})\in\mbox{$\boldsymbol{c}$}^{*}C^{\infty}(\mathbb{T}^{2}), we have

limnϕn(f,g,h)=2𝕋2f~𝑑g~dh~.\lim_{n\to\infty}\phi_{n}(f,g,h)=2\int_{\mathbb{T}^{2}}\tilde{f}d\tilde{g}\wedge d\tilde{h}.

Moreover, the limit only depends on f,g,h𝐜C(𝕋2)f,g,h\in\mbox{$\boldsymbol{c}$}^{*}C^{\infty}(\mathbb{T}^{2}). Therefore, the functional ϕ(f,g,h)=limnϕn(f,g,h)\displaystyle\phi(f,g,h)=\lim_{n\to\infty}\phi_{n}(f,g,h) is a cyclic 22-cocycle on 𝐜C(𝕋2)\mbox{$\boldsymbol{c}$}^{*}C^{\infty}(\mathbb{T}^{2}).

Proof.

Let n\boxplus_{n} be the set of equilateral cells that consist of 2n2^{n}-fold subdivision of II. By the first-order approximation, for g~,h~C(𝕋2)\tilde{g},\tilde{h}\in C^{\infty}(\mathbb{T}^{2}) we have

g~0,1\displaystyle\tilde{g}_{0,1} =g~(1)g~(0)=2ng~x(0)+o(2n)(n),\displaystyle=\tilde{g}(1)-\tilde{g}(0)=2^{-n}\tilde{g}_{x}(0)+o(2^{-n})\quad(n\to\infty),
h~1,2\displaystyle\tilde{h}_{1,2} =h~(2)h~(1)=2nh~y(0)+o(2n)(n),\displaystyle=\tilde{h}(2)-\tilde{h}(1)=2^{-n}\tilde{h}_{y}(0)+o(2^{-n})\quad(n\to\infty),
g~0,3\displaystyle\tilde{g}_{0,3} =g~(3)g~(0)=2ng~y(0)+o(2n)(n),\displaystyle=\tilde{g}(3)-\tilde{g}(0)=2^{-n}\tilde{g}_{y}(0)+o(2^{-n})\quad(n\to\infty),
h~3,2\displaystyle\tilde{h}_{3,2} =h~(2)h~(3)=2nh~x(0)+o(2n)(n).\displaystyle=\tilde{h}(2)-\tilde{h}(3)=2^{-n}\tilde{h}_{x}(0)+o(2^{-n})\quad(n\to\infty).

Apply the above equations to f~(0)(g~0,1h~1,2g~0,3h~3,2)\tilde{f}(0)(\tilde{g}_{0,1}\tilde{h}_{1,2}-\tilde{g}_{0,3}\tilde{h}_{3,2}) in Lemma 3.1, and we get

f~(0)(g~0,1h~1,2g~0,3h~3,2)=4nh~(0)(g~x(0)h~y(0)g~y(0)h~x(0))+o(4n)(n).\tilde{f}(0)(\tilde{g}_{0,1}\tilde{h}_{1,2}-\tilde{g}_{0,3}\tilde{h}_{3,2})=4^{-n}\tilde{h}(0)(\tilde{g}_{x}(0)\tilde{h}_{y}(0)-\tilde{g}_{y}(0)\tilde{h}_{x}(0))+o(4^{-n})\quad(n\to\infty).

Therefore, we have

limnnh~(0)(g~0,1h~1,2g~0,3h~3,2)\displaystyle\phantom{=}\lim_{n\to\infty}\sum_{\square\in\boxplus_{n}}\tilde{h}(0)(\tilde{g}_{0,1}\tilde{h}_{1,2}-\tilde{g}_{0,3}\tilde{h}_{3,2})
=limn4nnh~(0)(g~x(0)h~y(0)g~y(0)h~x(0))+4no(4n)\displaystyle=\lim_{n\to\infty}4^{-n}\sum_{\square\in\boxplus_{n}}\tilde{h}(0)(\tilde{g}_{x}(0)\tilde{h}_{y}(0)-\tilde{g}_{y}(0)\tilde{h}_{x}(0))+4^{n}o(4^{-n})
=𝕋2f~(g~xh~yg~yh~x)𝑑xdy\displaystyle=\int_{\mathbb{T}^{2}}\tilde{f}(\tilde{g}_{x}\tilde{h}_{y}-\tilde{g}_{y}\tilde{h}_{x})dx\wedge dy
=𝕋2f~𝑑g~dh~.\displaystyle=\int_{\mathbb{T}^{2}}\tilde{f}d\tilde{g}\wedge d\tilde{h}.

Similary, we obtain

limnnf~(2)(g~2,3h~3,0g~2,1h~1,0)\displaystyle\lim_{n\to\infty}\sum_{\square\in\boxplus_{n}}\tilde{f}(2)(\tilde{g}_{2,3}\tilde{h}_{3,0}-\tilde{g}_{2,1}\tilde{h}_{1,0}) =limnnf~(1)(g~1,0h~0,3g~1,2h~2,3)\displaystyle=-\lim_{n\to\infty}\sum_{\square\in\boxplus_{n}}\tilde{f}(1)(\tilde{g}_{1,0}\tilde{h}_{0,3}-\tilde{g}_{1,2}\tilde{h}_{2,3})
=limnnf~(3)(g~3,2h~2,1g~3,0h~0,1)\displaystyle=-\lim_{n\to\infty}\sum_{\square\in\boxplus_{n}}\tilde{f}(3)(\tilde{g}_{3,2}\tilde{h}_{2,1}-\tilde{g}_{3,0}\tilde{h}_{0,1}) =𝕋2f~𝑑g~dh~.\displaystyle=\int_{\mathbb{T}^{2}}\tilde{f}d\tilde{g}\wedge d\tilde{h}.

Hence, by Lemma 3.1, we get

limnϕn(f,g,h)\displaystyle\lim_{n\to\infty}\phi_{n}(f,g,h) =12limn𝒔S×nTr(f~𝒄[F,g~𝒄][F,h~𝒄]M)\displaystyle=\frac{1}{2}\lim_{n\rightarrow\infty}\sum_{\mbox{$\boldsymbol{s}$}\in S^{\times n}}\operatorname{Tr}\left(\tilde{f}\circ\mbox{$\boldsymbol{c}$}\Bigl{[}F,\tilde{g}\circ\mbox{$\boldsymbol{c}$}\Bigr{]}\left[F,\tilde{h}\circ\mbox{$\boldsymbol{c}$}\right]M\right)
=12limnnTr(f~[F,g~][F,h~]M)\displaystyle=\frac{1}{2}\lim_{n\rightarrow\infty}\sum_{\square\in\boxplus_{n}}\operatorname{Tr}\left(\tilde{f}[F,\tilde{g}][F,\tilde{h}]M\right)
=12limnn(f~(0)(g~0,1h~1,2g~0,3h~3,2)\displaystyle=\frac{1}{2}\lim_{n\rightarrow\infty}\sum_{\square\in\boxplus_{n}}\left(\tilde{f}(0)(\tilde{g}_{0,1}\tilde{h}_{1,2}-\tilde{g}_{0,3}\tilde{h}_{3,2})\right.
+f~(2)(g~2,3h~3,0g~2,1h~1,0)\displaystyle\hskip 68.2866pt+\tilde{f}(2)(\tilde{g}_{2,3}\tilde{h}_{3,0}-\tilde{g}_{2,1}\tilde{h}_{1,0})
f~(1)(g~1,0h~0,3g~1,2h~2,3)\displaystyle\hskip 68.2866pt-\tilde{f}(1)(\tilde{g}_{1,0}\tilde{h}_{0,3}-\tilde{g}_{1,2}\tilde{h}_{2,3})
f~(3)(g~3,2h~2,1g~3,0h~0,1))\displaystyle\hskip 68.2866pt\left.-\tilde{f}(3)(\tilde{g}_{3,2}\tilde{h}_{2,1}-\tilde{g}_{3,0}\tilde{h}_{0,1})\right)
=2𝕋2f~𝑑g~dh~.\displaystyle=2\int_{\mathbb{T}^{2}}\tilde{f}d\tilde{g}\wedge d\tilde{h}.

Assume that f=𝒄(f~)=𝒄(f~)f=\mbox{$\boldsymbol{c}$}^{*}(\tilde{f})=\mbox{$\boldsymbol{c}$}^{*}(\tilde{f}^{\prime}), g=𝒄(g~)=𝒄(g~)g=\mbox{$\boldsymbol{c}$}^{*}(\tilde{g})=\mbox{$\boldsymbol{c}$}^{*}(\tilde{g}^{\prime}) and h=𝒄(h~)=𝒄(h~)h=\mbox{$\boldsymbol{c}$}^{*}(\tilde{h})=\mbox{$\boldsymbol{c}$}^{*}(\tilde{h}^{\prime}). In general, if 𝒄(p~)=𝒄(q~)\mbox{$\boldsymbol{c}$}^{*}(\tilde{p})=\mbox{$\boldsymbol{c}$}^{*}(\tilde{q}) for p~,q~C(𝕋2)\tilde{p},\tilde{q}\in C(\mathbb{T}^{2}), then 𝒄n(p~)=𝒄n(q~)\mbox{$\boldsymbol{c}$}^{*}_{n}(\tilde{p})=\mbox{$\boldsymbol{c}$}^{*}_{n}(\tilde{q}) on the boundary of CDnCD_{n} for any nn\in\mathbb{N}. Therefore, for every nn\in\mathbb{N}, we have

𝒔S×nTr(f~𝒄[F,g~𝒄][F,h~𝒄]M)\displaystyle\sum_{\mbox{$\boldsymbol{s}$}\in S^{\times n}}\operatorname{Tr}\left(\tilde{f}\circ\mbox{$\boldsymbol{c}$}\Bigl{[}F,\tilde{g}\circ\mbox{$\boldsymbol{c}$}\Bigr{]}\left[F,\tilde{h}\circ\mbox{$\boldsymbol{c}$}\right]M\right) =𝒔S×nTr(f~𝒄[F,g~𝒄][F,h~𝒄]M)\displaystyle=\sum_{\mbox{$\boldsymbol{s}$}\in S^{\times n}}\operatorname{Tr}\left(\tilde{f}^{\prime}\circ\mbox{$\boldsymbol{c}$}\Bigl{[}F,\tilde{g}^{\prime}\circ\mbox{$\boldsymbol{c}$}\Bigr{]}\left[F,\tilde{h}^{\prime}\circ\mbox{$\boldsymbol{c}$}\right]M\right)

Thus as nn\to\infty, the value ϕ(f,g,h)\phi(f,g,h) depends only on f,g,h𝒄C(𝕋2)f,g,h\in\mbox{$\boldsymbol{c}$}^{*}C^{\infty}(\mathbb{T}^{2}). ∎

Corollary 3.5.

Given [p]K0(𝐜(C(𝕋2)))[p]\in K_{0}(\mbox{$\boldsymbol{c}$}^{*}(C^{\infty}(\mathbb{T}^{2}))) written as p=e𝐜:CDMN()p=e\circ\mbox{$\boldsymbol{c}$}:CD\rightarrow M_{N}(\mathbb{C}), the Connes’ pairing of the cyclic 22-cocycle [ϕ][\phi] with [p][p] is expressed as follows:

[ϕ],[p]=1πi𝕋2Tr(e(de)2).\langle[\phi],[p]\rangle=\dfrac{1}{\pi i}\int_{\mathbb{T}^{2}}\operatorname{Tr}(e(de)^{2}).

Acknowledgments

Seto was supported by JSPS KAKENHI Grant Number 21K13795.

References

  • [1] R. Baer. Abelian groups without elements of finite order. Duke Math. J., 3(1):68–122, 1937.
  • [2] A. Connes. Noncommutative differential geometry. Inst. Hautes Études Sci. Publ. Math., 62:257–360, 1985.
  • [3] T. Maruyama and T. Seto. A combinatorial fredholm module on self-similar sets built on nn-cubes, 2019. arXiv:1912.05832.
  • [4] H. Moriyoshi. On a cyclic volume cocycle in fractal geometry. talk on “Further developement of Atiyah-Singer Index Theory”, 2013 (based on joint work with T. Natsume).